Peptide-Based Nanoparticle Exhibiting

Jun 28, 2017 - Abstract | Full Text HTML | PDF w/ Links | Hi-Res PDF · A pH-Responsive Yolk-Like Nanoplatform for Tumor Targeted Dual-Mode Magnetic Re...
4 downloads 14 Views 1MB Size
Article

Self-Assembled DNA/Peptide-Based Nanoparticle Exhibiting Synergistic Enzymatic Activity Qing Liu, Hui Wang, Xinghua Shi, Zhen-Gang Wang, and Baoquan Ding ACS Nano, Just Accepted Manuscript • DOI: 10.1021/acsnano.7b03195 • Publication Date (Web): 28 Jun 2017 Downloaded from http://pubs.acs.org on June 28, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Nano is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

Self-Assembled DNA/Peptide-Based Nanoparticle Exhibiting Synergistic Enzymatic Activity Qing Liu†‡#, Hui Wang†#, Xinghua Shi†‡*, Zhen-Gang Wang†*, Baoquan Ding†‡* †

CAS Key Laboratory of Nanosystem and Hierarchial Fabrication, CAS Center for Excellence in

Nanoscience, National Center for Nanoscience and Technology, Beijing 100190 (P.R.China) ‡

University of Chinese Academy of Sciences, Beijing 100049, P. R. China

#

These authors contributed equally. * Address correspondence to: [email protected], [email protected], [email protected]

ACS Paragon Plus Environment

1

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 30

ABSTRACT. Designing enzyme-mimicking active sites in artificial systems is key to achieving catalytic efficiencies rivaling those of natural enzymes and can provide valuable insight in the understanding of the natural evolution of enzymes. Here, we report the design of a catalytic hemin-containing nanoparticle with self-assembled guanine-rich nucleic acid/histidine-rich peptide components, that mimicks the active site and peroxidative activity of hemoproteins. The chemical complementarities between the folded nucleic acid and peptide enable the spatial arrangement of essential elements in the active site and effective activation of hemin. As a result, remarkable synergistic effects of nucleic acid and peptide on the catalytic performances were observed. The turnover number of peroxide reached the order of that of natural peroxidase, and the catalytic efficiency is comparable to that of myoglobin. These results have implications in the precise design of supramolecular enzymes mimetics, particularly those with hierarchical active sites. The assemblies we describe here may also resemble an intermediate in the evolution of contemporary enzymes from the catalytic RNA of primitive cells.

KEYWORDS. DNA, peptide, hemin, peroxidase mimics, self-assembly

ACS Paragon Plus Environment

2

Page 3 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

Enzymes play vital roles in cellular functions by catalyzing biochemical reactions. High catalytic efficiency is largely attributed to the close cooperation among reactive elements in the active sites, which have been sufficiently screened and permutated by nature. Enzyme-inspired supramolecular catalysis, which harnesses multiple weak intermolecular interactions to assemble catalytic species, has attracted considerable interest and shown great potential in medicinal and industrial biotransformation applications.1-5 However, it has been a challenge to structurally and chemically reproduce the active sites of natural enzymes in supramolecular assemblies. Nucleic acid nanotechnology has provided a toolbox for the precise design of supromolecular structures and functions.6-9 With an in vitro selection (SELEX, Systematic Evolution of Ligands by Exponential enrichment) approach, catalytic nucleic acids have been produced to possess oxidase,10 peroxidase,11,12 deoxyribozyme13,14 or ligase activities.15 These activities of nucleic acid structures support the hypothesis of “RNA world”,16 where RNA is capable of selfreplication and catalyzing key metabolic pathways in primitive cells. However, the nucleic acidbased catalysts exhibit low catalytic efficiencies and limited catalytic reactions, which are likely attributable to the lack of key functional groups or proper geometrical arrangement at the active sites. SELEX may improve the catalytic properties by optimizing the screened sequences, which is usually a time-consuming and unpredictable process. Furthermore, the well-defined and sophisticated structures of enzymes illustrate that evolution and the increased complexity of life may be dominated by a hybrid catalytic system that evolved from RNA. However, the chemical forms of such a hybrid system remain unknown. A self-assembled peptide nanostructure may have served as an oasis for the prebiotic chemical evolution of RNA.17 Interestingly, the effective catalysis of some contemporary RNA-based enzymes is achieved via RNA/protein collaboration, such as in the CRISPER-Cas9 system18 or aminoacyl-tRNA synthetase19 in which case the

ACS Paragon Plus Environment

3

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 30

peptide motif is highly conserved. DNA and RNA exhibit exquisite shape and functional group complementarity to self-assembled peptides to form DNA or RNA-binding protein complexes.20,21 This latter fact has inspired us to design self-assembled nucleic acid/peptide hybrid systems with synergistic and tunable catalytic behaviors. Nature has evolved active sites in enzymes where the distribution of essential functional groups follows precise spatial configuration for effective activation of the catalytic center, such as in ferrochelatase,22 formate dehydrogenase23 , heme-copper oxidase24 and hemoproteins.25 Iron (III) protoporphyrin IX (hemin) is present in hemoproteins as a cofactor. In horseradish peroxidase, a common hemoprotein, hemin as a cofactor is activated by H2O2 via the cooperation among a proximal His ligand, a distal His and a distal Arg residue to form the primary active species, Compound I25. This species accepts the electrons from the reducing substrate, thereby leading to a variety of fundamental biological functions. The chemical groups and their spatial arrangement (Scheme S1) are prerequisites for the high catalytic efficiency. To create a similar active site facilitating the formation of compound I, here we designed a hybrid system that assembled a guanine-rich nucleic acid and a histidine-rich peptide into a ligand environment for hemin, as shown in Scheme 1. The coordination, catalytic properties and the activation mechanism of the hemin were explored. The nucleic acid strand can fold into a high-order structure composed of base-stacked quartets, which leads to hemin binding and enhancement of the peroxidase activities of hemin.26,27 The His rich peptide provides amino acid residues for complexation with and activation of hemin.28,29 Our work reveals that the nucleic acid and peptide ligands had compositional and structural complementarities to form the enzyme-like hierarchical active site for synergistic catalytic activity, which could be facilely tailored via the design of the nucleic acid and peptide molecules. Moreover, the hybrid system exhibited higher

ACS Paragon Plus Environment

4

Page 5 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

catalytic efficiency and turnover number than previously reported nucleic acid-based peroxidase mimetics.30,31 These hybrid assemblies may even be comparable in catalysis kinetics to natural hemeproteins.

Scheme 1. Self-assembly of guanine-rich nucleic acid, His-rich peptide and the cofactor hemin into the peroxidase-mimicking nanoparticles. The designed nucleic acid and peptide components exhibit structurally and chemically complementary characteristics that enable the creation of the peroxidase-mimicking hierarchical active site in the artificial enzyme. RESULTS AND DISCUSSION We first modeled a single-molecule-level rigid docking using the Autodock program to show that the 32 His-containing peptide molecule (H32) can approach the folded, guanine-rich DNAzyme-I (GGGTAGGGCGGGTTGGG; DzI) from several directions (Figure S1-3). This preliminary approach indicated the shape complementarity between these two species. A molecular dynamics (MD) simulation was subsequently carried out for the two most stable DNA/peptide conformations (Figure S4) to identify van der Waals forces as the driving force for DzI/H32 self-assembly. In buffer conditions, the self-aggregation of H32 into spherical

ACS Paragon Plus Environment

5

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 30

nanoparticles was observed (Figure S5). This behavior is attributed to the intermolecular association of β-sheets of H32 as revealed by the CD results (Figure S6). The H32/DzI assembly disturbed the conformation of the DzI G-quadruplex, without significantly altering the secondary structure of H32 or nanoparticle morphologies. The Zetapotential results (Figure S7a) indicate the effective interactions of DzI with H32, which increased the negativity of Zeta potential of H32-based nanoparticles significantly. The binding mode of H32 to DzI was studied by competitive binding assay (Figure S7b). Hoechst 33258 fluorescent dyes, which were classic groove binders of DNA, were lighted up upon bound to DzI.32,33 The addition of H32 to the DzI/Hoechst complex reduced the fluorescent intensity by 36%, illustrating the aggregates of His-rich peptide probably competitively bound to G-quadruplex of DzI through groove binding, a common mode for DNA-protein recognition.34 The scanning transmission electron microscopy (STEM) imaging and elemental mapping results revealed the successful complexation of hemin with H32 and DzI (Figures 1A and 1B) by quantifying the elemental contribution of P (only found in DNA) and Fe (only found in hemin) related to the total signal. These data illustrate the uniform distribution of DzI and hemin in the nanospheres. (For more TEM images, see Figure S8.) DzI and H32 can both form coordination bonds with hemin (2.26 Å and 2.14 Å for G- and Hishemin iron bond lengths, respectively). To understand the organization of the hybrid selfassemblies it is essential to identify the interactions between the hemin iron and the selfassembled DzI and H32 components. We investigated this using UV-vis spectroscopy (Figure 1C). The H32/hemin assembly exhibited a Soret band at 414 nm, and a pronounced band at 530 nm and shoulder at 570 nm. The spectral features are attributed to the imidazole→iron chargetransfer transition and the low-spin six-coordinated species35,36 with two strong-field ligands (one

ACS Paragon Plus Environment

6

Page 7 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

ligand on the sixth coordination site, presumably His residue).37,38 The DzI/hemin assembly showed a Soret band at 403 nm and charge-transfer transitions at 626 nm and 504 nm with a weak absorbance at 530 nm, which are characteristic spectral features of high-spin hexacoordinated species11,39,40 with a strong-field and weak-field ligand (water) at the axial hemin coordination positions.37,41 In the DzI/H32/hemin assembly, the Soret peak of hemin underwent a blue shift from 414 nm and leveled off at 408 nm (a red shift compared to DzI/hemin), as the DzI:H32 ratio increased. Meanwhile, the absorbance bands of DzI-complexed hemin iron in the visible region (450-700 nm) turned more evident, which indicates the preferential coordination of hemin iron to the nucleobase ligand of DzI. (For the dependence of DzI/hemin and DzI/H32/hemin spectra on DzI concentrations, see Figure S9.)

Enhanced charge-transfer

transitions of hemin iron were observed, which is indicative of the stronger interactions of the dπ orbitals of the metal with the π orbitals of the pyrrole rings.42 This could be attributed to a decrease in the symmetry of the hemin moiety, which may arise from the packing forces of the peptide chains43 and the cooperative stacking effect of His residues and nucleobases on the local electric fields.44,45

ACS Paragon Plus Environment

7

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 30

Figure 1. (A) TEM images of the DzI/H32/hemin assembly. Scale bar: 500 nm. (B) Elementresolved STEM images of the self-assembled DzI/H32/hemin nanoparticles. Scale bar: 200 nm. [Hemin]: 100 nM. [H32]: 4 µM. [DzI]: 1 µM. (C) UV-vis spectra of free hemin (black), H32hemin (blue), DzI (red) and DzI/H32 (green)-complexed hemin. [Hemin]: 500 nM. [H32]: 4 µM. [DzI]: 3 µM. Figure S10 shows the common hemin location for all of the lowest energy docking structures. The hemin stacks on the G-quartet and the exocyclic amine of the G9 base (GGGTAGGGCGGGTTGGG) lies close to and axial to the hemin iron. Within the DzI/H32/hemin assembly, the active site, in which a guanine ligand and a distal His are distributed to either side of hemin, may be rationally proposed. The neighboring His residues from the H32 aggregates can provide or accept hydrogen bonds in a role similar to the distal Arg

ACS Paragon Plus Environment

8

Page 9 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

in natural peroxidases.25,46 The heteromolecular self-assembly rigidifies the spatial location of the His residues that are catalytically important for natural peroxidase and stabilizes hemin via hydrogen bonding between His and the water ligand. These features underscore the complementary roles of the DNA and peptide components in the formation of a 3D hierarchical active site and the activation of peroxidase-mimicking catalysis by hemin. H2O2-mediated catalyzed oxidization of ABTS2- (2,2’-azino-bis(3-ethylbenzothiazoline-6sulfonic acid)) was studied by monitoring the time-dependent absorbance changes of the reaction product, ABTS•+, at 415 nm (the extinction coefficient is 31100 M-1 cm-1) (Figure S11). The initial catalytic velocity (Vi) and the maximum catalytic conversion efficiency of ABTS2- (CABTS) were used to evaluate the catalytic performance. Figure 2A shows that hemin encapsulated within the DzI/H32 assembly showed enhanced catalytic performance compared to hemin with DzI or H32 alone, indicating a cooperation of the peptide and DNA in the catalytic activity of hemin. At high concentrations of H2O2, an inactivation of the catalytic assemblies was observed, which resulted in a stronger synergistic effect in particular at 20 mM H2O2. This effect is also observed in natural peroxidase subjected to excessive H2O2, and was ascribed to the formation of a reactive ferroporphyrin species followed by hemin destruction.47 The catalytic performances of the assemblies revealed a dependence on the DzI:H32 ratio (Figures 2B and S11). At 20 mM H2O2, Vi increased and leveled off at 2 µM DzI, and CABTS reached a maximum at 1.5 µM DzI. The performance of the DzI/hemin assembly was proportional to the DzI concentration within the studied range. At an appropriate DzI/H32 ratio, CABTS and Vi increased by up to 2 and 2.5 fold, respectively. Even at a DzI concentration as low as 20 nM, 5-fold lower than hemin, the assembly resulted in enhanced reaction conversion.

ACS Paragon Plus Environment

9

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 30

Figure 2 The dependence of catalytic performance indicators Vi and CABTS of the artificial enzymes on (A) H2O2 concentration. [DzI]: 2 µM, [H32]: 4 µM. (B) DzI concentration. [H2O2]: 20 mM. [hemin]: 100 nM. [H32]: 4 µM. It was suggested by Brown48 that the amino acid residues adjacent to the methylene bridges of the porphyrin ring protect peroxidase active sites against H2O2 oxidization. The end-stacking of hemin at the terminals of the folded DzI results in an accessible surface at the catalytic center and causes hemin to be susceptible to H2O2,49 which accounts for the higher Vi and faster degradation rate than H32-complexed hemin. The His-rich peptide protected hemin against oxidative degradation, a feature that is apparent in the reactions catalyzed by the H32-containing assembly.

ACS Paragon Plus Environment

10

Page 11 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

The degradation of hemin at different H2O2 concentrations was also investigated by recording the time-dependent UV-vis spectral changes of hemin (Figure S13). The self-assembly of DNA with peptide decreased the H2O2-induced destruction of the DNA-complexed hemin. Table 1. Apparent kinetic parameters with respect to catalyzing reduction of H2O2 with ABTS2as the reducing substrate. kcat

Km

kcat / Km

(s-1)

(mM)

(s-1 mM-1)

DzI/hemin [b]

0.32

2.58

0.124

H32/hemin[b]

0.18

2.51

0.0717

DzI/H32/hemin [b]

0.436

1.71

0.254

DzI/H32/hemin [c]

3.01

5.04

0.597

Catalytic complexes[a]

[a] [DzI]: 2 µM,[H32]: 4 µM. [b] 100 nM hemin. [c] 1.5 nM hemin

To investigate the enzymatic kinetics, Lineweaver-Burk plots of the initial catalytic velocity versus H2O2 concentration were constructed for the three artificial enzymes containing 100 nM hemin (Figure S14). The apparent kinetic parameters for H2O2 reduction in the presence of ABTS2- were calculated using the Michaelis-Menten equation (Table 1). The kcat value, which indicates an enzyme’s turnover number (TON), and kcat/Km, which reflects the catalytic efficiency, are both consistent with a synergistic effect of DzI and H32 ligands in creating a highly active artificial peroxidase. The superiority of the DzI/H32/hemin assembly was further verified when the ratio of hemin to the DzI/H32 components was decreased. At a hemin concentration of 10 nM (Figures S15a and S15b), the optimal Vi and CABTS for the DzI/H32/hemin complex were 3.5-fold and 4-fold those of H32/hemin, and 34-fold and 84-fold those of DzI/hemin, respectively. At 1.5 nM hemin

ACS Paragon Plus Environment

11

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 30

(Figures S15c and S15d), the optimal Vi and CABTS for the DzI/H32/hemin were 10-fold and 20fold those of H32/hemin, and no activity for DzI/hemin was observed. Even in the presence of 0.2 nM DzI (a far low DzI : H32 ratio), an enhancement in the substrate conversion was observed. Moreover, as shown in Table 1, the kcat value (3.01 s-1) for the assembly containing 1.5 nM hemin was almost one order higher than that that containing 100 nM hemin (0.436 s-1) (Figure S16) and approached that of 1.5 nM natural horseradish peroxidase (approximately 18.2 s-1). The catalytic efficiency (kcat/Km) of 1.5 nM hemin-contained complex was almost identical to myoglobin50 and the kcat value was almost two orders higher. It is noteworthy that the kcat/Km and kcat values were much higher than previously reported nucleic acid-based peroxidase mimetics generated by in vitro selection or modulation.30,31 Together these data show that the His residues can promote the activation of DzI-complexed hemin. The investigation of the effect of H32 concentration on the catalytic properties of the hybrid nanoparticles, with fixed DzI/H32 ratio (1/4) and hemin concentration (1.5 nM), indicate that the catalytic activity increased and leveled off at ca. 12 µM H32 (Figure S17). The catalytic activities of the self-assemblies are tailorable by varying the sequences of DNA or the length of the peptide ligands. We designed another four guanine-rich DNA strands adapted from DzI sequence: DzII to DzV, which can all fold into a G-quadruplex (for the sequences, see Figure S18). As shown in Figure 3A and Figure S19a, the guanine-rich DNA/H32/hemin assemblies exhibited activity following the order of the corresponding DNA/hemin. This is ascribed to the effectiveness of the nucleoligand coordination to hemin (Figure S19b and c). We then altered the length of the His-rich peptide component to contain 1, 9 or 20 residues (i.e., designated as H1, H9 and H20). The optimal Vi and CABTS values of the H20 complex was between those of H9 and H32 (Figure 3B; for the time-dependent courses, see Figure S20). The

ACS Paragon Plus Environment

12

Page 13 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

CD results (Figure S21) and the competitive binding assay (Figure S7b) indicate that the longer peptide interacts more strongly with DzI, which may result in a more effective synergistic activation of hemin. No synergistic effect was observed in the H1/DzI system in the catalytic activity assay or UV-vis spectra (Figure S22).

Figure 3. The effect of (A) guanine-rich DNA sequences (DzI-DzV) (1.5 nM hemin) and (B) His-rich peptide length (H9, H20 and H32) (100 nM hemin) on the catalytic performance of the DNA/peptide/hemin assemblies. [Guanine-rich DNA]: 1.5 µM, [His-rich peptides]: 4µM. The effect of the component species was also investigated using guanine-free DNA strands or His-free peptides. As shown in Figure S23, the use of non-cognate DNA resulted in the inactivation of the H32-complexed hemin. Similarly, the assembly with the non-cognate peptide decreased the catalytic activity of the DzI-complexed hemin (Figure S24). These results confirm the importance of the key functional groups in the nucleic acid and peptide ligands in the cooperative activation of hemin-catalyzed reactions. In addition to ABTS2-, the synergistic effects in the catalyzed oxidization of TMB (3,3’,5,5’tetramethylbenzidine), phenol, homovanillic acid (a major catecholamine metabolite), dopamine

ACS Paragon Plus Environment

13

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 30

(a catechol derivative as a neurotransmitter) and pyrogallol were also observed for DzI/H32/hemin nanoparticles (Figure S25-S29). The kinetic parameters for the reduction of H2O2 in the presence of pyrogallol were indicated in Table S1, which also showed that the kinetic parameters of other reported hemin-based catalytic system. It is found that the kcat value for hemin complexed with DzI/H32 was higher than that for hemin conjugated with cyclodextrin,51encapsulated in supramolecular hydrogel which contained Phe and His residues.51 Particularly, at low ratio of hemin to the DzI/H32 hybrid, the kcat and kcat/Km values were higher than hemin/graphene conjugates,52 and were more closely to HRP.53

This indicated the

superiority of our catalytic system based on heteromolecular self-assembly. Furthermore, in addition to DNA, the assembly of guanine-rich RNA with H32 and hemin exhibited synergistic catalytic activity (Figure S30). Using the lowest energy mechanism of H2O2 activation of horseradish peroxidase46 as a paradigm, a model was constructed for the artificial active site to describe the roles of the peptide and nucleic acid components in the formation of reactive intermediate compound I based on the B3LYP54,55 functional of Density Functional Theory (DFT; Figure 4).56-59 The coordinating G9 nucleobase and three distal His groups (one Hisα and two neighboring Hisβ residues) are arranged around the hemin (state i). H2O2 (Hα-Oα-Oβ-Hβ) is held by the Hisα/Hisβ via double hydrogen bonds, N (Hisα)···Hβ-Oβ and N-H (Hisβ)···Oβ, and a Fe-Oα bond (state ii). This is followed by the abstraction of a distal proton by Hisα from H2O2 and a flip of the Fe-Oα-Hα-Oβ moiety to the FeOβ-Oα-Hα complex, which is stabilized by hydrogen bonding between Hisα (H+)/Hisβ and Hβ-OαOβ groups (state iii). Then, the Hβ-Oα group on the distal side of Fe is reprotonated by Hisα (H+) and the heterolytic cleavage of the Oα-Oβ bond leads to the formation of the Compound I analogue (state iv), which is facilitated by a negative charge on the exocyclic amine of the G9

ACS Paragon Plus Environment

14

Page 15 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

base and a positive charge on Hisα (H+). The bond lengths are in line with effective coordination and the hydrogen bonds.

In this model, the His groups act as the acid-base catalyst and

hydrogen-bond acceptors/donors, the guanine-rich DNA (or RNA) with the stacked quartets coordinate with and stabilize the hemin iron, and the heteromolecular self-assembly enhances the G9-Fe coordination (as revealed by the spectra in Figure 1C).

Figure 4. The QM models of the self-assembled active site involved in the formation of Compound I (state iv) based on DFT/B3LYP calculations. The N atoms of two Hisβ residues and the porphyrin ring are fixed in the geometry optimization process. These results show the cooperative effect of the three His and the G9 coordination group on the H2O2 activation of hemin. The electron-withdrawing effect of the polar imidazole groups60,61 adjacent to the porphyrin ring may facilitate the electron transfer from the reducing aromatic substrate to Compound I, which is consistent with the higher turnover rate of ABTS2- (kcat: 0.036 s-1) for the DzI/H32/hemin assembly (0.018 s-1 for DzI/hemin and 0.016 s-1 for H32/hemin in the presence of 100 nM hemin) (For the time-dependent absorbance changes, see Figure S31). The peptide chains also provide protection for the hemin against H2O2-induced degradation. The

ACS Paragon Plus Environment

15

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 30

experimental and theoretical results substantiate the attribution of the synergistic peroxidative catalysis to the formation of an enzyme-like active site. We also examined the effect of species assembly order, and found an optimal catalytic performance for DzI + H32 + hemin (i.e., the addition of hemin to DzI/H32 hybrid; Figure S32). This result reveals that the pre-assembly of DzI with H32 enables a proper distribution of the reactive elements and full catalytic cooperation between the DNA and peptide components. CONCLUSIONS We have described the self-assembly of nucleic acid with peptide in the construction of enzyme-mimicking catalytic nanoparticles with tailorable activities. Employing hemin as the cofactor, the self-assembled nanoparticles exhibited significantly enhanced peroxidasemimicking activity in the oxidization of a variety of reducing substrates, compared to solely nucleic acid- or peptide-based assemblies.

The synergistic catalysis was attributed to the

effective cooperation between the nucleic acid and peptide components which possess complementary chemical and structural characteristics. Importantly, the spatial arrangement and functions of the reactive groups in the artificial active site resemble those in natural peroxidase. The activity of the hybrid nanoparticles is highly dependent on the guanine-rich DNA that acted as a molecular scaffold for hemin, the His-rich peptide that provided activating groups and hemin that performed the rodox catalysis. This system may serve as spectrophotometric or spectrofluorimetric biosensors for i) sensitive in vitro detection of DNA (e.g. telomeric ordered DNA62), based on that only 0.2 nM guanine-rich DNA could result in enhancement of the substrate conversion (Figure S15d); ii) probing the activity of proteinases (e.g. matrix metallopteinases (MMPs) or caspases) that cleaved the specific-sequence peptide bridging oligomeric His residues and altered the catalytic activities.

The structural and functional

ACS Paragon Plus Environment

16

Page 17 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

complexity and diversity of our catalytic system can be enhanced through appropriate design of the core components or cofactors for effective catalytic bond cleavage or formation reactions. Our work may also provide a laboratory model for a self-assembled prebiotic intermediate between the RNA-based catalytic system and contemporary enzymes. MATERIALS AND METHODS Materials All peptides with purity level above 98% were purchased from Shanghai ZiYu Biotech Co.,Ltd. All DNA oligonucleotides (purified with dual PAGE) were purchased from Invitrogen Life Technologies (Shanghai, China). RNA (with 2’Ome modification) was purchased from Shanghai GenePharma Co., Ltd, and purified with HPLC. Hemin, HEPES/HEPES sodium salt, Na2HPO4/NaH2PO4

salts,

H2O2,

horseradish

peroxidase

(HRP),

2,2’-Azino-bis(3-

ethylbenzothiazoline-6-sulfonic acid) diammonium salt (ABTS2−), Hoechst 33258 dye, 3,3',5,5'Tetramethylbenzidine (TMB), phenol, o-dianisidine, homovanillic acid, pyrogallol and DMSO were purchased from Sigma-Aldrich. Activity assay As-received lyophilized DNA or peptides were dissolved in deionized water to make 100 µM or 400 µM stock solution, and stored at -20oC. Repetitive freeze-thawing operations should be avoided. Hemin was dissolved in DMSO to make a 100 nM-10µM stock solution, stored at room temperature. DNA and peptide with required concentrations were mixed with hemin for 20 min. The assembly time was determined by the CD signals of DNA (the conformation was disturbed by the peptide) and UV-vis absorbance of hemin, ensuring the assembly efficiency of the complex.

Freshly prepared H2O2 and the reducing substrates were added to the solution

containing the assembled complexes, and the time-dependent absorbance change at the reactant

ACS Paragon Plus Environment

17

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 30

or the product was recorded, which were used to calculate the initial catalytic velocity and the maximum conversion efficiency. Each measurement was repeated four times. Characterizations The samples for TEM imaging were prepared by placing 5 µL of the sample solution on a glow-discharged carbon-coated grid (400 meshes, Ted Pella), followed by the wicking away of the unbound sample and solution evaporation. The TEM characterization was conducted using a Hitachi H-7700 microscope operating at 80 kV. The CD spectra were measured by a Jasco J-810 spectropolarimeter. The Ultraviolet-visible absorption spectra were recorded using a UV-VIS spectrophotometer

(UV-2450,

Shimadzu).

The

fluorescence

was

measured

with

a

spectrofluorometry (F-4500, Hitachi, Tokyo, Japan) (for Hoechst 33258 dye, λex= 346 nm). Scanning transmission electron microscopy (STEM) and the elemental mapping were performed in FEI Tecnai F20 using a high-angle annular dark field (HAADF) detector, coupled with an energy dispersive X-ray spectrometer (EDX). Zeta-potential measurements were performed using a Malvern Zeta sizer Nano ZS instrument Theoretical Simulations Multi-scale computational methods (quantum-chemical calculations, molecular docking and molecular dynamic simulation) were applied to understand the activation center by exploring the structure, dynamics and interaction of DNA DzI, peptide H32 and hemin. The DNA model was built on the basis of Bcl-2 (pdb:2f8U)63 by retaining the conformation of G-quadruplexes and replacing the loop nucleobases. Fe(III) protoporphyrin IX (pdb:2QSP)64 from bovine haemoglobin were used as hemin models. All the MD simulations were performed using the Gromacs 5.0 package65,66 combined with the AMBER03 force field67 and TIP4P solvent model.68 Na+ was used as counter ions to neutralize the systems and the temperature and pressure were

ACS Paragon Plus Environment

18

Page 19 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

maintained using V-rescale thermostat69 and isothermal-isobaric ensemble.70 Particle Mesh Ewald (PME)71 was employed to deal with the long-range electrostatic interactions. The peptide and DNA conformations from the MD equilibration were used for molecular docking. All the molecular docking simulations were carried out with Lamarckian Genetic Algorithm (LGA) using Auto-Dock 4.2.6.72 A big grid box size of 126 × 126 × 126 points with a large spacing of 0.753 Å (DNA & peptide) and 0.375 Å (DNA & Heme) between the grid points was implemented and the grid box is big enough to cover the entire surface of the DNA. The ones with lowest binding energy were selected for the detailed analysis and further MD studies. The structures of the activation center model were optimized with Gaussian 09 program and the stationary points were confirmed to be minima by vibrational analysis. All the calculations were carried out at the B3LYP/6-31G*54,55 level of Density Functional Theory (DFT).56,58,59. For the more detailed computional methods, see Note S1-S4.

ASSOCIATED CONTENT Supporting Information. Details for computational simulations, Scheme 1, Figure S1-S31 and the corresponding captions, DNA and RNA sequences. AUTHOR INFORMATION Corresponding Author *[email protected], [email protected], [email protected] Author Contributions

ACS Paragon Plus Environment

19

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 30

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. #These

authors contributed equally.

ACKNOWLEDGMENT The authors are grateful for the financial support from National Basic Research Programs of China (2016YFA0201601), National Science Foundation China (21273052, 21573051, 11422215, 11272327 and 11672079), Beijing Municipal Science & Technology Commission (No. Z161100000116036), CAS Interdisciplinary Innovation Team, National Program for Support of Top-notch Young Professionals, Youth Innovation Promotion Association CAS, CAS Key Laboratory of Nanosystem and Hierarchical Fabrication, REFERENCES 1. Meeuwissen, J.; Reek, J. N. H. Supramolecular Catalysis Beyond Enzyme Mimics. Nat. Chem. 2010, 2, 615-621. 2. Dong, Z. Y.; Wang, Y. G.; Yin, Y. Z.; Liu, J. Q. Supramolecular Enzyme Mimics by SelfAssembly. Curr. Opin. Colloid Interface Sci. 2011, 16, 451-458. 3. Wiester, M. J.; Ulmann, P. A.; Mirkin, C. A. Enzyme Mimics Based upon Supramolecular Coordination Chemistry. Angew. Chem. Int. Ed. 2011, 50, 114-137. 4. Wu, L. Z.; Chen, B.; Li, Z. J.; Tung, C. H. Enhancement of the Efficiency of Photocatalytic Reduction of Protons to Hydrogen via Molecular Assembly. Acc. Chem. Res. 2014, 47, 21772185.

ACS Paragon Plus Environment

20

Page 21 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

5. Lin, Y. H.; Wu, L.; Huang, Y. Y.; Ren, J. S.; Qu, X. G. Positional Assembly of Hemin and Gold Nanoparticles in Graphene-Mesoporous Silica Nanohybrids for Tandem Catalysis. Chem. Sci. 2015, 6, 1272-1276. 6. Seeman, N. C. DNA in a Material World. Nature 2003, 421, 427-431. 7. Fu, J. L.; Liu, M. H.; Liu, Y.; Yan, H. Spatially-Interactive Biomolecular Networks Organized by Nucleic Acid Nanostructures. Acc. Chem. Res. 2012, 45, 1215-1226. 8. Wilner, O. I.; Willner, I. Functionalized DNA Nanostructures. Chem. Rev. 2012, 112, 25282556. 9. Albinsson, B.; Hannestad, J. K.; Borjesson, K. Functionalized DNA Nanostructures for Light Harvesting and Charge Separation. Coord. Chem. Rev. 2012, 256, 2399-2413. 10. Golub, E.; Freeman, R.; Willner, I. A Hemin/G-Quadruplex Acts as an NADH Oxidase and NADH Peroxidase Mimicking DNAzyme. Angew. Chem. Int. Ed. 2011, 50, 11710-11714. 11. Travascio, P.; Li, Y. F.; Sen, D. DNA-Enhanced Peroxidase Activity of a DNA AptamerHemin Complex. Chem. Biol. 1998, 5, 505-517. 12. Golub, E.; Albada, H. B.; Liao, W. C.; Biniuri, Y.; Willner, I. Nucleoapzymes: Hemin/GQuadruplex DNAzyme–Aptamer Binding Site Conjugates with Superior Enzyme-Like Catalytic Functions. J. Am. Chem. Soc. 2016, 138, 164-172. 13. Liu, J. W.; Cao, Z. H.; Lu, Y. Functional Nucleic Acid Sensors. Chem. Rev. 2009, 109, 1948-1998.

ACS Paragon Plus Environment

21

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 30

14. Silverman, S. K. Catalyic DNA: Scope, Applications, and Biochemistry of Deoxyribozymes. Trends Biochem. Sci. 2016, 41, 595-609. 15. Cuenoud, B.; Szostak, J. W. A DNA Metalloenzyme with DNA-Ligase Activity. Nature 1995, 375, 611-614. 16. Orgel, L. E. Prebiotic Chemistry and the Origin of the RNA World. Crit. Rev. Biochem. Mol. Biol. 2004, 39, 99-123. 17. Carny, O.; Gazit, E. A Model for the Role of Short Self-Assembled Peptides in the Very Early Stages of the Origin of Life. FASEB J. 2005, 19, 1051-1055. 18. Shalem, O.; Sanjana, N. E.; Zhang, F. High-Throughput Functional Genomics Using CRISPER-Cas9. Nat. Rev. Genet. 2015, 16, 299-311. 19. Passioura, T.; Suga, H. Reprogramming the Genetic Code in Vitro. Trends Biochem. Sci. 2014, 39, 400-408. 20. Rohloff, J. C.; Gelinas, A. D.; Jarvis, T. C.; Ochsner, U. A.; Schneider, D. J.; Gold, L.; Janjic, N. Nucleic Acid Ligands with Protein-Like Side Chains: Modified Aptamers and Their Use as Diagnostic and Therapeutic Agents. Mol. Ther.-Nucl. Acids 2014, 3, e201. 21. Davies, D. R.; Gelinas, A. D.; Zhang, C.; Rohloff, J. C.; Carter, J. D.; O'Connell, D.; Waugh, S. M.; Wolk, S. K.; Mayfield, W. S.; Burgin, A. B.; Edwards, T. E.; Stewart, L. J.; Gold, L.; Janjic, N.; Jarvis, T. C. Unique Motifs and Hydrophobic Interactions Shape the Binding of Modified DNA Ligands to Protein Targets. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 1997119976.

ACS Paragon Plus Environment

22

Page 23 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

22. Al-Karadaghi, S.; Hansson, M.; Nikonov, S.; Jonsson, B.; Hederstedt, L. Crystal Structure of Ferrochelatase: The Terminal Enzyme in Heme Biosynthesis. Structure 1997, 5, 1501-1510. 23. Shi, J. F.; Jiang, Y. J.; Jiang, Z. Y.; Wang, X. Y.; Wang, X. L.; Zhang, S. H.; Han, P. P.; Yang, C. Enzymatic Conversion of Carbon Dioxide. Chem. Soc. Rev. 2015, 44, 5981-6000. 24. Bhagi-Damodaran, A.; Michael, M. A.; Zhu, Q. H.; Reed, J.; Sandoval, B. A.; Mirts, E. N.; Chakraborty, S.; Moënne-Loccoz, P.; Zhang, Y.; Lu, Y. Why Copper Is Preferred over Iron for Oxygen Activation and Reduction in Haem-Copper Oxidases. Nat. Chem. 2017, 9, 257-263. 25. Veitch, N. C. Horseradish Peroxidase: A Modern View of a Classic Enzyme. Phytochemistry 2004, 65, 249-259. 26. Sen, D.; Poon, L. C. H. RNA and DNA Complexes with Hemin [Fe(III) Heme] Are Efficient Peroxidases and Peroxygenases: How Do They Do It and What Does It Mean? Crit. Rev. Biochem. Mol. Biol. 2011, 46, 478-492. 27. Wang, Z. G.; Liu, Q.; Ding, B. Q. Shape-Controlled Nanofabrication of Conducting Polymer on Planar DNA Templates. Chem. Mater. 2014, 26, 3364-3367. 28. Gao, Y.; Zhao, F.; Wang, Q. G.; Zhang, Y.; Xu, B. Small Peptide Nanofibers as the Matrices of Molecular Hydrogels for Mimicking Enzymes and Enhancing the Activity of Enzymes. Chem. Soc. Rev. 2010, 39, 3425-3433. 29. Singh, N.; Tena-Solsona, M.; Miravet, J. F.; Escuder, B. Towards Supramolecular Catalysis with Small Self-Assembled Peptides. Isr. J. Chem. 2015, 55, 711-723.

ACS Paragon Plus Environment

23

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 30

30. Mao, X. H.; Simon, A. J.; Pei, H.; Shi, J. Y.; Li, J.; Huang, Q.; Plaxco, K. W.; Fan, C. H. Activity Modulation and Allosteric Control of a Scaffolded DNAzyme Using a Dynamic DNA Nanostructure. Chem. Sci. 2016, 7, 1200-1204. 31. Zhu, L.; Li, C.; Zhu, Z.; Liu, D. W.; Zou, Y.; Wang, C. M.; Fu, H.; Yang, C. J. In Vitro Selection of Highly Efficient G-Quadruplex-Based DNAzymes. Anal. Chem. 2012, 84, 83838390. 32. Neidle, S. DNA Minor-Groove Recognition by Small Molecules. Nat. Prod. Rep. 2001, 18, 291-309. 33. Wang, J. S.; Zhao, C. Q.; Zhao, A. D.; Li, M.; Ren, J. S.; Qu, X. G. New Insights in Amyloid Beta Interactions with Human Telomerase. J. Am. Chem. Soc. 2015, 137, 1213-1219. 34. Tateno, M.; Yamasaki, K.; Amano, N.; Kakinuma, J.; Koike, H.; Allen, M. D.; Suzuki, M. DNA Recognition by β-Sheets. Biopolymers 1997, 44, 335-359. 35. Monzani, E.; Bonafe, B.; Fallarini, A.; Redaelli, C.; Casella, L.; Minchiotti, L.; Galliano, M. Enzymatic Properties of Human Hemalbumin. Biochim. Biophys. Acta-Protein Struct. Molec. Enzym. 2001, 1547, 302-312. 36. Shantha, P. K.; Saini, G. S. S.; Thanga, H. H.; Verma, A. L. Photoreduction of Iron Protoporphyrin IX Chloride in Non-Ionic Triton X-100 Micelle Studied by Electronic Absorption and Resonance Raman Spectroscopy. J. Raman Spectrosc. 2001, 32, 159-165. 37. Babul, J.; Stellwagen, E. Participation of the Protein Ligands in the Folding of Cytochrome c. Biochemistry 1972, 11, 1195-1200.

ACS Paragon Plus Environment

24

Page 25 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

38. Lombardi, A.; Nastri, F.; Pavone, V. Peptide-Based Heme-Protein Models. Chem. Rev. 2001, 101, 3165-3189. 39. Yonetani, T.; Anni, H. Yeast Cytochrome c Peroxidase. Coordination and Spin States of Heme Prosthetic Group. J. Biol. Chem. 1987, 262, 9547-9554. 40. Ozaki, S.; Hara, I.; Matsui, T.; Watanabe, Y. Molecular Engineering of Myoglobin: The Improvement of Oxidation Activity by Replacing Phe-43 with Tryptophan. Biochemistry 2001, 40, 1044-1052. 41. Stellwagen, E.; Babul, J. Stabilization of the Globular Structure of Ferricytochrome c by Chloride in Acidic Solvents. Biochemistry 1975, 14, 5135-5140. 42. Zelet, B.; Kaposi, A. D.; Nucci, N. V.; Sharp, K. A.; Dalosto, S. D.; Wright, W. W.; Vanderkooi, J. M. Water Channel of Horseradish Peroxidase Studied by the Charge-Transfer Absorption Band of Ferric Heme. J. Phys. Chem. B 2004, 108, 10317-10324. 43. Jentzen, W.; Ma, J. G.; Shelnutt, J. A. Conservation of the Conformation of the Porphyrin Macrocycle in Hemoproteins. Biophys. J. 1998, 72, 753-763. 44. Manas, E. S.; Vanderkooi, J. M.; Sharp, K. A. The Effects of Protein Environment on the Low Temperature Electronic Spectroscopy of Cytochrome c and Microperoxidase-11. J. Phys. Chem. B 1999, 103, 6334-6348. 45. Prabhu, N. V.; Dalosto, S. D.; Sharp, K. A.; Wright, W. W.; Vanderkooi, J. M. Optical Spectra of Fe (II) Cytochrome c Interpreted Using Molecular Dynamics Simulations and Quantum Mechanical Calculations. J. Phys. Chem. B 2002, 106, 5561-5571.

ACS Paragon Plus Environment

25

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 30

46. Derat, E.; Shaik, S. The Poulos-Kraut Mechanism of Compound I Formation in Horseradish Peroxidase: A QM/MM Study. J. Phys. Chem. B 2006, 110, 10526-10533. 47. Valderrama, B.; Ayala, M.; Vazquez-Duhalt, R. Suicide Inactivation of Peroxidases and the Challenge of Engineering More Robust Enzymes. Chem. Biol. 2002, 9, 555-565. 48. Brown, S. B. Stereospecific Heam Cleavage. Biochem. J. 1976, 159, 23-27. 49. Yang, X. J.; Fang, C. L.; Mei, H. C.; Chang, T. J.; Cao, Z. H.; Shangguan, D. H. Characterization of G-Quadruplex/Hemin Peroxidase: Substrate Specificity and Inactivation Kinetics. Chem.-Eur. J. 2011, 17, 14475-14484. 50. Wan, L. L.; Twitchett, M. B.; Eltis, L. D.; Mauk, A. G.; Smith, M. In Vitro Evolution of Horse Heart Myoglobin to Increase Peroxidase Activity. Proc. Natl. Acad. Sci. U. S. A. 1998, 95, 12825-12831. 51. Wang, Q. G.; Yang, Z. M.; Zhang, X. Q.; Xiao, X. D.; Chang, C. K.; Xu, B. A Supramolecular-Hydrogel-Encapsulated Hemin as an Artificial Enzyme to Mimic Peroxidase. Angew. Chem. Int. Ed. 2007, 46, 4285-4289. 52. Xue, T.; Jiang, S.; Qu, Y. Q.; Su, Q.; Cheng, R.; Dubin, S.; Chiu, C. Y.; Kaner, R.; Huang, Y.; Duan, X. F. Graphene-Supported Hemin as a Highly Active Biomimetic Oxidation Catalyst. Angew. Chem. Int. Ed. 2012, 51, 3822-3825. 53. Yamaguchi, H.; Tsubouchi, K.; Kawaguchi, K.; Horita, E.; Harada, A. Peroxidase Activity of Cationic Metalloporphyrin-Antibody Complexes. Chem.-Eur. J. 2004, 10, 6179-6186. 54. Becke, A. D. Density-Functional Thermochemistry. 3. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648-5652.

ACS Paragon Plus Environment

26

Page 27 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

55. Lee, C. T.; Yang, W. T.; Parr, R. G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785-789. 56. Delley, B.; Wrinn, M.; Luthi, H. P. Binding Energies, Molecular Structures, and Vibrational Frequencies of Transition Metal Carbonyls Using Density Functional Theory with Gradient Corrections. J. Chem. Phys. 1994, 100, 5785-5791. 57. Jonas, V.; Thiel, W. Theoretical Study of the Vibrational Spectra of the Transition Metal Carbonyls M(Co)6 [M=Cr, Mo, W], M(Co)5 [M=Fe, Ru, Os], and M(Co)4 [M=Ni, Pd, Pt]. J. Chem. Phys. 1995, 102, 8474-8484. 58. Ehlers, A. W.; Frenking, G. Structures and Bond Energies of the Transition Metal Hexacarbonyls M(Co)6 (M=Cr, Mo, W). A Theoretical Study. J. Am. Chem. Soc. 1994, 116, 1514-1520. 59. Li, J.; Schreckenbach, G.; Ziegler, T. A Reassessment of the First Metal-Carbonyl Dissociation Energy in M(Co)4 (M=Ni, Pd, Pt), M(Co)5 (M=Fe, Ru, Os), and M(Co)6 (M=Cr, Mo, W) by a Quasi-Relativistic Density-Functional Method. J. Am. Chem. Soc. 1995, 117, 486494. 60. Kulhanek, J.; Bures, F. Imidazole as a Parent π-Conjugated Backbone in Charge-Transfer Chromophores. Beilstein J. Org. Chem. 2012, 8, 25-49. 61. Bures, F. Fundamental Aspects of Property Tuning in Push-Pull Molecules. RSC Adv. 2014, 4, 58826-58851.

ACS Paragon Plus Environment

27

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 30

62. Zhao, C. Q.; Wu, L.; Ren, J. S.; Xu, Y.; Qu, X. G. Targeting Human Telomeric HigherOrder DNA: Dimeric G-Quadruplex Units Serve as Preferred Binding Site. J. Am. Chem. Soc. 2013, 135, 18786-18789. 63. Dai, J. X.; Chen, D.; Jones, R. A.; Hurley, L. H.; Yang, D. Z. NMR Solution Structure of the Major G-Quadruplex Structure Formed in the Human BCL2 Promoter Region. Nucleic Acids Res. 2006, 34, 5133-5144. 64. Aranda, R.; Cai, H.; Worley, C. E.; Levin, E. J.; Li, R.; Olson, J. S.; Phillips, G. N.; Richards, M. P. Structural Analysis of Fish versus Mammalian Hemoglobins: Effect of the Heme Pocket Environment on Autooxidation and Hemin Loss. Proteins 2009, 75, 217-230. 65. Pronk, S.; Pall, S.; Schulz, R.; Larsson, P.; Bjelkmar, P.; Apostolov, R.; Shirts, M. R.; Smith, J. C.; Kasson, P. M.; van der Spoel, D.; Hess, B.; Lindahl, E. Gromacs 4.5: A HighThroughput and Highly Parallel Open Source Molecular Simulation Toolkit. Bioinformatics 2013, 29, 845-854. 66. Hess, B.; Kutzner, C.; van der Spoel, D.; Lindahl, E. Gromacs 4: Algorithms for Highly Efficient, Load-Balanced, and Scalable Molecular Simulation. J. Chem. Theory Comput. 2008, 4, 435-447. 67. Duan, Y.; Wu, C.; Chowdhury, S.; Lee, M. C.; Xiong, G. M.; Zhang, W.; Yang, R.; Cieplak, P.; Luo, R.; Lee, T.; Caldwell, J.; Wang, J. M.; Kollman, P. A Point-Charge Force Field for Molecular Mechanics Simulations of Proteins Based on Condensed-Phase Quantum Mechanical Calculations. J. Comput. Chem. 2003, 24, 1999-2012.

ACS Paragon Plus Environment

28

Page 29 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

68. Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R. W.; Klein, M. L. Comparison of Simple Potential Functions for Simulating Liquid Water. J. Chem. Phys. 1983, 79, 926-935. 69. Bussi, G.; Donadio, D.; Parrinello, M. Canonical Sampling through Velocity Rescaling. J. Chem. Phys. 2007, 126, 014101. 70. Parrinello, M.; Rahman, A. Polymorphic Transitions in Single-Crystals: A New Molecular Dynamics Method. J. Appl. Phys. 1981, 52, 7182-7190. 71. Essmann, U.; Perera, L.; Berkowitz, M. L.; Darden, T.; Lee, H.; Pedersen, L. G. A Smooth Particle Mesh Ewald Method. J. Chem. Phys. 1995, 103, 8577-8593. 72. Morris, G. M.; Huey, R.; Lindstrom, W.; Sanner, M. F.; Belew, R. K.; Goodsell, D. S.; Olson, A. J. Autodock4 and Autodocktools4: Automated Docking with Selective Receptor Flexibility. J. Comput. Chem. 2009, 30, 2785-2791.

ACS Paragon Plus Environment

29

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 30

For Table of Contents Only

ACS Paragon Plus Environment

30